Está en la página 1de 11

Materials Research.

2012; 15(3): 372-382

2012

DDOI: 10.1590/S1516-14392012005000035

Chemical and Topographic Analysis of Treated Surfaces of Five Different


Commercial Dental Titanium Implants
Bruno Ramos Chrcanovic*, Alexsander Ribeiro Pedrosa, Maximiliano Delany Martins
Laboratrio de Nanoscopia LabNano, Centro de Desenvolvimento da Tecnologia Nuclear CDTN,
Comisso Nacional de Energia Nuclear CNEN, Av. Presidente Antnio Carlos, 6627,
Campus da UFMG, Pampulha, CEP 31270-901, Belo Horizonte, MG, Brazil
Received: August 22, 2011; Revised: February 7, 2012

We present a detailed investigation of the surface characteristics of five commercial titanium


implants with different surface finishing (double acid etching, anodization and incorporation of
Ca/P, acid etching and deposition of Ca/P, hydroxyapatite-blasting, acid etching and Ca/P-blasting)
produced by five different manufacturers. A set of experimental techniques were employed to study
the surface chemical composition and morphology: XPS, XRD, SEM, EDS, and AFM. According to
the implat manufacturers, the addition of Ca and P at the implant surface is a main feature of these
implants (except the double acid etched implant, which was included for comparative purpose).
However, the results showed a great discrepancy on the final amount of these elements on the implant
surface, which suggests a different effectiveness of the employed surface finishing methods to fix those
elements on the implant surface. Our results show that only the method used by the manufacturer of
hydroxyapatite-blasting surface finished implants was efficient to produce a hydroxyapatite coating.
This group also showed the highest roughness parameters.
Keywords: dental implants, surface modification, chemical analysis, titanium

1. Introduction
Oral rehabilitation by means of endosseous dental
implants is an essential issue in clinical practice. Local
bone quality and systemic implications on the oral healing
condition have a direct role in the success of dental implant
therapy1.
Since the finding of the osseointegration concept, the
characteristics of the interface between bone and implant,
and possible ways to improve it, have been of particular
interest in dental and orthopedic implant research.
Osseointegration is defined experimentally as the close
contact between bone and implant material in histological
sections and, in clinical terms, as the stability and ankylosis
of an implant in bone2.
Upon the placement of an implant into a surgical site,
there is a cascade of molecular and cellular processes that
provide for new bone growth and differentiation along
the biomaterial surface3. One means to improve implant
success is through methods to increase the amount of bone
contact along the body of the implant. While it may seem
obvious that increased surface roughness of implants leads
to greater success it is not clear what aspect of roughness
is advantageous4.
Already by the beginning of the 1980s, surface structure
was identified as one of the six main factors particularly
important for implant incorporation into bone5, a statement
that has been confirmed in later published research. Faster
and stronger bone formation may confer better stability
*e-mail: brunochrcanovic@hotmail.com

during the healing process, thus allowing more rapid loading


of the implant. The aim of rough surfaces is not only to
increase the interlock between the implant and bone, but
also to improve the bone-healing process4.
An increasing number of surface modifications are
introduced and despite a majority of studies comparing
machined surfaces with new rough surfaces, it is not
clear whether, in general, one surface modification is
better than another. To further add to the confusion, not
only surface topography is changed with many techniques
but also surface chemistry and altered topography
commonly results in a change in the chemistry and vice
versa4.
The fact that the surface properties of titanium and of
other biomaterials show considerable variation depending
on the type of surface treatment emphasizes the necessity of
surface characterization of the materials used in biological
experiments. A systematic approach towards understanding
the relationship between material properties and biological
response is of course crucially dependent on knowledge
about the surface properties of the investigated materials6.
Well-defined surface characterization provides a scientific
basis for a better understanding of the effects of the implant
surface on the biological response. Moreover, a more detailed
study on the characterization of dental implants commercially
available in Brazil is lacking. Thus, the purpose of the present
study was to investigate the superficial chemical composition
and surface topography of five titanium dental implants
commercially available in Brazil.

2012; 15(3)

Chemical and Topographic Analysis of Treated Surfaces of Five Different Commercial Dental Titanium Implants

2. Materials and Methods


Five screw-shaped implants from different manufacturers
and different surface modifications were chosen for analysis.
For each type of implant two specimens were purchased
in the market. The implants were all received in their
original sterile packaging and only opened at the start of
the investigation. They were carefully handled in order to
prevent contamination during further manipulations. They
were classified into five groups [model, manufacturer]:
(1) double acid etching (StylusTM, SIN, So Paulo, Brazil),
(2) anodization and electrochemical incorporation of Ca/P
(Vulcano ActivesTM, Conexo, So Paulo, Brazil), (3) acid
etching and deposition of Ca/P (NanotiteTM, Biomet 3i, Palm
Beach Gardens, USA), (4) hydroxyapatite-blasting (HAp
ceramedTM, Ceramed, Lisboa, Portugal), and (5) acid etching
and Ca/P-blasting (OsseanTM, Intra-lock, Boca Raton,
USA). A detailed explanation of the surface modification
methods was not available to all sample groups and was
not provided by the manufacturers. From the information
available, it can be said that the implants from group 3 have
a dual-acid surface etching treatment followed by deposition
of nanometer scale crystals of calcium phosphate (CaP).
The surface treatment is by discrete crystalline deposition
on the surface, not a plasma sprayed coating. The coating
process for group 4 is performed by thermal spraying of
hydroxyapatite (HA) on the titanium surface. The HA is
heated to a molten or semi-molten state within the projection
gun, and projected on a surface previously prepared. Then
the material solidifies, giving rise to multiple layers that
form the coating. The implants of group 5 are previously
acid-etched and then blasted with particles of calcium and
phosphate in clean atmosphere.
The five different titanium-treated surfaces were
examined by scanning electron microscopy - SEM (JEOL,
model JSM-840A, Tokyo, Japan). The SE mode with an
acceleration voltage of 25kV was selected for SEM analysis
and the vacuum pressure was maintained below 1105 Torr.
The load current (LC) was approximately 85A. For a
direct comparison of the surface morphology, the same
magnification was used for all investigated implants.
In order to obtain quantitative analysis of the implant
roughness, atomic force microscopy (AFM - NTegra
Aura, NT-MDT, Moscow, Russia) of the implant samples
was performed. AFM images were acquired in air using
semicontact mode with a NSG 01 sharpened gold-coated
silicon tip (nominal spring constant of 2.510N/m and
nominal resonance frequency of 110200kHz, NT-MDT).
The scanning area for the measurements was 2020m2.
The images obtained by AFM were characterized by 2nd
order extraction filter, using the software Image Analysis
2.1.2 (NTMDT, Moscow, Russia). The seven surface
roughness parameters calculated by the software were
evaluated (Sy, Sz, Sa, Sq, Ssk, and Ska). The mean value
and standard deviation of these parameters was obtained
from ten satisfactory scans of each group, from random
sites of the flat surface of the groove from the apex of the
implants.The surface chemical composition was analyzed by
energy dispersive x-ray spectroscopy - EDS (JEOL, model
JXA-8900RJ, Tokyo, Japan). Two regions of the implants
were evaluated: the platform, in order to analyze the type

373

of titanium substrate (Ti/cp or Ti6Al4V), since no surface


modification was present at this region, and the flat surface
of the groove from the apex of the implants.
The surface chemical composition was also analyzed
by X-ray photoelectron spectroscopy (XPS) using MgK
radiation (1253.6eV) and a CLAM2 Electron Analyzer
(VG Microtech, East Sussex, UK). XPS analysis provides more
superficial information (few nanometers deep) compared to
EDS, which was use to better understand the chemical profile
introduced by the different modification methods. The x-ray
beam was directed to middle of the threads of the implants. The
base pressure was below 2.01010 mbar during spectra data
acquisition. Survey XPS data was collected from 0 to 1100eV
of binding energy and acquired with constant pass energy of
100eV. High resolution XPS spectra were obtained at Ti2p,
O1s, C1s, Ca2p, P2p, N1s, and Si2p, when present, with
constant pass energy of 50eV. XPS data were acquired after
30minutes surface cleaning by Ar+ sputtering. All binding
energies were referenced relative to the main hydrocarbon
peak (from residual hydrocarbon contamination) set at
285.0eV. The high resolution XPS peaks were analyzed using
GaussianLorentzian curve fitting and Shirley peak baseline
with the Xi Spectral Data Processor 32bit version 3.0 software
(XPS International, Los Altos, USA), in order to determine peak
positions, areas and full-width half maximums (FWHM). The
elemental surface composition was calculated by normalizing
the relative spectral peak areas (A) using relative sensitivity
factors S7, as illustrated in the equation:
Cx =

Ax
A
i (1)
S x i Si

where Cx is the atomic fraction of any constituent in the


sample surface.
X-ray diffraction measurements were carried out in a
D/Max Ultima X-ray diffractometer (Rigaku, Tokyo, Japan),
using Cu-K1 radiation without any filter or monochromator,
in the angle range of 10-90 (2) with a grazing incidence
of 3 degrees, so that diffraction can be made surface
sensitive. The results were analyzed in Search-Match
(Crystallographica, Oxford, United Kingdom) and Fullprof
(CNRS, Gif sur Yvette, France) softwares. The diffraction
patterns for each compound used in the comparisons were
standardized by JCPDS-International Center for Diffraction
Data (PCPDFWIN v.2.01, ICCD, Newton Square, USA).
The indexation of patterns of crystal structures was
determined according to Cullity and Stock8. The indexing
is provided by the diffraction lines with a value of sin2,
which satisfies a certain equation for each crystal structure.
For a cubic crystal, the equation:
sin 2
2

(h + k + l )

2
4a2

(2)

is obtained by combining the Bragg law with the


plane-spacing equationfor the cubic system. Patterns of
hexagonal crystals can be determined by the following
equation:
sin 2 =

2
2

2 4 h + hk + k
l2

+ 2 (3)
2
4 3
a
c

374

Chrcanovic et al.

and the sin2 values in the tetragonal system must obey


the relation:
sin 2 =

2
4a

(h

+ k2 +

2
4c 2

Materials Research

distribution of spherical structures arranged like clusters


of grapes (see detail 10000). Group 5 implants were
mainly characterized with facets produced by blasting and
fine etching pits, with some smooth areas and other very
rough areas (Figure1e).

l 2 (4)

where h, k and l are the Miller indices, a and c are the lattice
parameters, is the wavelength of the X-ray beam, and is
the angle of diffraction in degrees.

3.2. EDS analysis


EDS-analysis of the surface showed titanium to be the
common element in screw thread and platform in all groups,
except for the screw thread region in group 4. There was high
purity of titanium in screw thread and platform in group 1,
and in the platform in groups 2 and 5. Calcium, phosphor
and oxygen were the additional elements found on the
screw thread in group 2. Ca, P, O and C were found in the
screw thread of group 4; Ti was not observed. The presence
of titanium, aluminum, and vanadium was observed in
screw thread and platform of group 3 and in the platform
of group 4. Calcium, phosphor and titanium were found in
the screw thread of group 5.

3. Results
3.1. SEM analysis
Figures1a to 1e, revealed characteristic differences at
the microscopic level according to the surface modification
methods used for the implant samples as measured by
SEM. Group 1 implants showed surface topography
subtracted by acid baths, almost like a needle-like elevation
structure (Figure1a). Pore size was 3.0-6.0m, sometimes
elongated to 10m. Group 2 implants showed a porous
structure induced by breakdown phenomena during
anodic oxidation like craters of volcanoes (Figure1b).
Pore size was 0.5-3.0m. Group 3 implants also showed
a needle-like elevation structure (Figure1c), as observed
in group 1, but the pore size was smaller (0.5-2.0m).
Group 4 implants showed a less homogeneous surface
structure than the other groups (Figure1d). The surface
appears to be thermally melted by the spray. There was a

3.3. AFM analysis


The qualitative and quantitative surface topography
demonstrated different degrees of roughness. Implants of the
groups 2, 3, and 5 exhibited similar roughness parameters.
Groups 1 and 4 exhibited a rougher surface than the other
three groups, as demonstrated by higher Sy, Sz, Sq, and Sa
parameters (Table1).

10 m

10 m

(a)

10 m

(b)

(c)

1 m

10 m
(d)

10 m
(e)

Figure1. SEM pictures of groups 1a); 2b); 3c); 4d); and 5e) (original magnification 3000).

2012; 15(3)

Chemical and Topographic Analysis of Treated Surfaces of Five Different Commercial Dental Titanium Implants

375

Table1. Mean values (SD) of tridimensional roughness parameters as determined by AFM.


Group

Sy
(m)

Sz
(m)

Average-mean
height (m)

Sa
(m)

Sq
(m)

Ssk

Sku

3.80 1.00

1.90 0.49

1.88 0.69

0.59 0.17

0.78 0.19

0.06 0.27

0.39 0.41

2
3
4
5

2.84 0.98
2.79 0.66
4.33 1.01
2.96 0.34

1.42 0.49
1.40 0.33
2.16 0.50
1.47 0.16

1.45 0.60
1.33 0.39
2.19 0.48
1.19 0.13

0.40 0.19
0.36 0.09
0.59 0.15
0.40 0.02

0.49 0.22
0.45 0.11
0.74 0.18
0.49 0.03

0.02 0.25
0.34 0.30
0.12 0.13
0.18 0.23

0.20 0.30
0.26 0.79
0.16 0.30
0.32 0.15

3.4. XPS analysis


The XPS survey spectra of as-received dental implant
showed major peaks at Ti2p, O1s and C1s (not presented
here), except for group 4, which did not show Ti on surface.
All elements except Ti and O decrease rapidly or disappear
after sputtering to a depth of a few nm, indicating that
the impurities are present predominantly at the outermost
surface. In all groups, the Ti peak intensities increased after
sputter cleaning (except for group 4) and the carbon peaks
disappeared. Surface elemental composition (% atomic
concentration), as measured by XPS analysis, and calculated
by Equation1, is presented in Table2.
High-resolution XPS spectra of Ti2p and O1s are
presented in Figure2. The O1s spectra showed main peaks
between 530.00-530.36eV. For sputter cleaned titanium
surfaces, this peak always dominates the O1s regions, as
was true for all the groups but group 4. O1s spectra were
deconvoluted to TiO2, (OH) or C-O, and Ti-OH or C=O.
The full width at half maximum (FWHM) of O1s (TiO2)
was smaller for group 5 implants and showed rather similar
features between the other groups. The O1s spectrum of
group 4 was deconvoluted to O2 (PO4, CO3 and OH), and
H2O,9 with higher FWHM (Table3).
The titanium spectra obtained by XPS were complex.
Titanium was found to be present in the metallic state (Ti)
and in the divalent (TiO), trivalent (TiO2) and tetravalent
(Ti2O3) oxide states. As already observed10, the peaks
corresponding to TiO2 are by far the most prominent ones
in the Ti2p region.
The Ti2p doublet peak, i.e., Ti2p1/2 and Ti2p3/2,
appeared between 464.50-464.74eV and 458.48-458.83eV
for all the implants. Ti2p1/2 were deconvoluted to
TiO2 at 464.580.09eV (MeanSD), Ti2O3+TiO
at 462.770.24eV, and Ti metal at 460.890.68eV
(Table4). Ti2p3/2 were deconvoluted to TiO 2 at
458.670.14eV, Ti2O3+TiO at 456.250.36eV, and Ti
metal at 453.221.81eV (Table5). The deconvolutions
were performed according to the other authors 11. The
FWHM was greater for group 2 implants for TiO2 peaks of
Ti2p and showed rather similar features between the other
groups (Tables4 and 5).
The C1s signals were curve-fitted based on three
different components. The deconvolution is straightforward
as well as the assignment to hydrocarbon (285.0eV), C-O
species (~286eV) and COO species (288-298eV)11. The
presence of carbon in different chemical environments is
typical for titanium (and other metal) oxide surfaces. This
is a consequence of the tendency of titanium surfaces to

readily adsorb organic components during fabrication (e.g.


lubricants) or from the ambient atmosphere during storage11.
Few additional elements were detected by XPS in very low
concentrations.

3.5. XRD analysis


Figures3a and e, show typical XRD pattern for the
groups 1 to 5, respectively. The diffraction peak positions
were calculated using the Equations2 to 4 and identified
by (hkl) Miller indices. The XRD measurements identified
crystal structures of Ti and titanium hydride (TiH2) in group 1,
Ti and anatase (TiO2) in group 2, Ti in group 3, Ti and HA
in group 4, and Ti in group 5. Ti and HA have hexagonal
close-packed crystal structure with lattice parameters
of a=0.295nm, c=0.468nm, and a=0.94344nm,
c=0.68815nm, respectively12,13, anatase has a tetragonal
structure with lattice parameters of a=0.3785nm,
c=0.9514nm,14 and TiH2 has a face-centered cubic crystal
structure with lattice parameter of a=0.4448nm15.

4. Discussion
The surface elemental composition and morphology of
five commercial titanium implants with different surface
finishing produced by five different manufacturers have
been analyzed using a complementary set of experimental
techniques (SEM, EDS, AFM, XPS, and XRD). However,
it is important to note that only two specimens per type
of implant were investigated. With such limited sample
investigation no conclusions can be drawn with regard
to the degree to which these samples are typical of the
manufacturers production, neither within the particular
batch nor between batches 11. On the other hand, the
experimental findings and the differences between the
various types of surfaces turned to be characteristic for
different areas of the same device as well as between the
two specimens investigated per implant type11. The present
results showed that surface chemistry and morphology of
different commercially available clinical titanium implants
differed due to surface modification techniques used during
manufacturing.

4.1. SEM analysis


Fine differences of the pits and facets were obvious
between the implants due to differences of the blasting
and etching processes. Each manufacturer makes use of a
specific treatment. The groups 1 and 3 showed fine structures
characteristically produced by chemical etching processes.
However, the two surfaces are clearly different, both as

376

Chrcanovic et al.

Table2. Surface elemental composition (% atomic concentration)


after 30minutes surface cleaning by Ar+ sputtering, as determined
by XPS (see text).
Element
Ti
O
Ca
P
Si

Implant groups
1

35.8
74.2
-

19.0
70.7
7.8
2.5
-

5.0
86.0
7.3
1.7
-

71.3
13.9
7.8
7.0

15.2
81.1
0.4
3.3
-

regards the dimensions of the surface microstructures as


well as the homogeneity. In the case of the groups 4 and 5,
a coarser microstructure as a consequence of a particle
blasting process can be observed and a finer etch-type
structure that is relatively constant in dimensions and fairly
regularly distributed across the surface was superimposed
on the coarser blasted structure. The electrochemical
oxidation process for group 2 implants changed not only
surface chemistry due to incorporation of anions from the
electrolyte, but also produced a porous structure of the
thick oxide layer.

4.2. EDS analysis


A similar pattern between the spectra of the platform in
groups 3 and 4 can be observed by the EDS analysis. In these
groups Ti, Al, and V are the substrate elements (suggesting
a Ti6Al4V alloy). The group 4 showed total covering of the
substrate, as none of these elements were observed in the
screw thread spectrum. Ti is the substrate in groups 1, 2,
and 5. Calcium was observed in groups 2, 3, 4 and 5 and
P in groups 2, 4, and 5. The different relative intensity of
calcium in these groups can be explained by the variation of
surface modification techniques. The expression of Ca and
total absence of P in group 3 indicates that the deposition
of CaP compound may have been insufficient in this sample
due to technical limitation.
The presence of oxygen only in groups 2 and 4 can be
explained by the composition of HA and phosphates used,
for its high reactivity with other elements of the environment
and by the process of alkaline passivation of the surface of
the implants made by these manufacturers. This element is
present in all groups in the form of oxides, as showed by the
XPS spectra. However, the microprobe has a relatively high
profile analysis, requiring the use of more surface sensitive
techniques such as XPS.

4.3. AFM analysis


The results of the present study demonstrate that
implants from the groups 1 and 4 exhibited a rougher surface
textures compared to the other groups. As the implant
designs were screw-shaped the AFM scans were achieved
from the flat surface of the apex of the implant, since the
AFM cannot be used on implant thread. However, a flat
apex area is often manufactured with another machining
technique than rest of the implant (often milled instead of
turned), and this will influence the final outcome. The value
may therefore not be representative for rest of the implant.

Materials Research

It is worth noticing that a large surface roughness


does not represent a problem for coated metallic implants
in biomedical applications, because roughness usually
ensures a better osseointegration, as compared to smoother
implants16. The advantage of small increase in roughness
was further demonstrated in rabbit femora and tibiae,
with commercially pure titanium (cp Ti) screws attracting
better bone response and greater torsional fixation when
the blasting texture increased from 1.10 to 1.45m
(Sa parameter)17. Thus, according to this previous study, an
average surface roughness of about 1.0-1.4m seems to be
most suitable for good bone-to-metal fixation, which was
not observed by any of the implants studied here. The use
of different roughness measurement devices, such as optical
profilometer18, confocal laser scanning profilometer19,
optical laser roughness tester instrument20 or AFM in the
present study may also have influenced the results. As
far as the quantification of surface characterization of
implants is concerned, it is at present an unresolved issue.
One concerning aspect is the resolution of profilometry.
The roughness of implant surfaces tends to be below the
resolution of conventional mechanical techniques (the
profilometric tip size, 35m). The laser profilometer
has proved to be a more suitable method to characterize
the roughness of implant surfaces with a laser-focused
spot of 1m and a measuring length of some millimeters.
The resolution of an AFM is even higher. The tip used in
the AFM in the present study has a tip curvature radius
of 10nm. This may also have influenced the results of
roughness parameters found in this study, which were much
lower than in previous studies that made measurements
using a profilometer18-20. The experimental methods that
uses larger scanning tips cannot resolve the finer features
of the etched surfaces4, and therefore higher values of the
surface roughness are expected. Moreover, higher roughness
values are found for longer measured distances, in line with
previous observations17. Thus, the smaller values observed
in the present study may have resulted from a smaller area
chosen measured by AFM for measurement (20x20m2)
in comparison with other studies that observed higher Sa
but with higher measured area (typically 240245m2),
and using larger scanning tips with a profilometer18,19. This
heterogeneity may have been caused by differences in
measurement methods.

4.4. XPS analysis


It is generally accepted that the outermost atomic layer
of the implant surface is a key factor in the osseointegration
process. As the cell-surface interaction takes place over
a few atomic distances, compositional modifications at
the atomic level on the implant surface can influence the
biocompatibility and the osseointegration prognosis of the
implant21. Therefore, it is very important to know accurately
the surface chemical composition of titanium dental implants
(as well as every other biomaterial).
Previous studies on the surface chemistry of machined,
turned and blasted implants have reported Ti, O and C as
major elements10,11,22,23. The chemistry-modified implants
have revealed the presence of Ca, P, Mg, S, F, and Na in
association with the used surface finishing methods10,22,24. In

2012; 15(3)

Chemical and Topographic Analysis of Treated Surfaces of Five Different Commercial Dental Titanium Implants

Figure2. Ti2p and O1s high-resolution XPS spectra for the 5 different groups (see text for experimental details).

377

378

Chrcanovic et al.

Materials Research

Table3. Binding energy and FWHM of O1s components, as determined after curve fitting of the XPS high resolution spectra (see text
for details).
Group
1
2
3
4
5

Peak (eV)
TiO2

(OH)s, C-O

TiOH, C=O

530.36
530.49
530.00

532.13
531.94
531.15

534.09
533.26
532.75

530.17

531.47

532.54

529.50

Table4. Binding energy and FWHM of Ti2p1/2 components, as


determined after curve fitting of the XPS high resolution spectra
(see text for details).
Group
1
2
3
5

FWHM
2

Peak (eV)
TiO2

Ti2O3 + TiO

Ti metal

464.50
464.74
464.45
464.57

462.50
463.08
462.63
462.94

460.50
461.65
460.81
461.48

531.60

H2O

533.00

2.10
1.90
2.00
4.40/2.60/4.40
1.60

Table5. Binding energy (EB) and FWHM of Ti2p3/2, as determined


after curve fitting of the XPS high resolution spectra (see text for
details).

FWHM

Group

2.20
2.61
2.10
2.00

1
2
3
5

the present study, the surface elements at the dental implants


mainly consisted of Ti, O and C. The spectra of all groups
were dominated by Ti and O due to the naturally formed
TiO2 layer. Concentrations of oxygen, carbon and titanium
on the surface of the implant groups showed differences that
can be attributed to the different processes they are subjected
during the fabrication process. The presence of silicon
found on the surface of group 4 implants is mainly due to
the contamination coming from manufacturing processes,
surface treatment and sterilization25.
A thin layer of titanium oxide was observed in all
implant groups, except group 4. The reason for this
absence was the fact that the XPS technique is very
sensitive to the most superficial region of a material, and
the implants from group 4 had a HA layer completely
covering the titanium, thick enough to avoid the
underneath substrate to be detected by XPS. There is
evidence that various properties of titanium oxide in
fact lead to different biological responses at several
levels, including biomolecular interactions, cellular
behavior, tissue response and biomechanical stability6.
This thin oxide film, naturally formed on a titanium
substrate, is presumably responsible for the excellent
biocompatibility of titanium implants due to a low level
of electronic conductivity26, a high corrosion resistance
and a thermodynamically stable state at physiological
pH values27. The differences in the percentage content
between the implants analyzed here were considered
to be related to the different methods of surface
modification performed by each manufacturer. The
speed and pressure of the instrumentation applied in the
surface modifications, surface temperature exposure to
air, lubricants and coolants used, all together influence
the nature of the implant surface 28 . The titanium
spontaneously forms a passivating oxide layer at ambient

PO4, CO3, OH

Peak (eV)

FWHM

TiO2

Ti2O3 + TiO

Ti metal

458.48
458.83
458.60
458.70

456.10
455.85
456.35
456.70

453.80
450.00
454.05
454.26

2.20
2.61
2.10
2.00

temperatures when exposed to air or water23. The XPS


high-resolution spectra are dominated by a doublet peak
at ~459 and ~464eV, which can be assigned to TiO2
(Ti2p3/2 and Ti2p1/2, respectively). A difference of
5.8 to 6.0eV in binding energies between the Ti2p3/2
and Ti 2p1/2 peaks of the doublet further corroborated
the presence of TiO 2 on all the surfaces 29. The XPS
measurements indicate that the titanium implants from the
group 4 coated by the HA-blasting process are completely
covered by HA, because no Ti peaks are observed in the
XPS spectra.
Minor contributions are frequently detected. These
spectral features may be fitted (by peak deconvolution) to
other oxidation states such as TiO, Ti2O3, Ti hydroxide,
Ti nitride, or Ti carbide6. In addition to TiO2, Ti2O3 and
TiO were also detected. The relative intensities of these
components are shown in Table 6. Hydrated water (Ti-OH)
and hydroxides (OH) were obtained by deconvolution of
Ti2p and O1s. Olefjord and Hansson22 demonstrated that
the outer oxide layers of Brnemark and IMZ implants were
TiO2, using an ESCA instrument, and this was confirmed by
the study of Sawaseetal.23. These layers were considered
to be formed during production and handling of implant
products23. At Ti2p, the peaks approximately corresponded
to the binding energy of TiO2. There were, however, minute
differences of less than 0.5eV between detected peaks and
binding energy of TiO2. Contamination by some carbonic
complexes adhering to the implant surface might cause the
minute charge shifts due to their insulating effects. The
observed chemical shift between the oxide peak and the
metal peak agrees well with the expected value for TiO230.
The oxygen envelope for all surfaces usually had at
least three components (sub-peaks). Sub-peaks at 530eV
were identified in groups 1, 2, 3, and 5. The presence of
hydroxides is indicated by the shape of the Ols peak for

2012; 15(3)

Chemical and Topographic Analysis of Treated Surfaces of Five Different Commercial Dental Titanium Implants

379

Table6. Relative peak areas of Ti2p and O1s components as determined by XPS analysis.
Group

Relative peak area (%)


Ti 2p

1
2
3
4
5

O 1s

TiO2

Ti2O3 + TiO

Ti metal

TiO2

62.1
64.9
63.7
84.0

13.6
24.6
15.3
10.2

24.2
10.5
20.9
5.9

61.8
60.8
77.0
56.1

(OH), C-O Ti-OH, C = O


33.9
30.4
17.5
28.8

4.3
8.8
5.4
15.1

O2

PO4, OH

H2O

42.3
-

36.5
-

21.2
-

Figure3. Typical XRD pattern for samples from group 1 to 5, a)to e)respectively.

these surfaces, which frequently shows spectral components


between 531 and 533eV, assigned to OH and H 20,
respectively31. These results thus show that surface oxides
on titanium implants should not generally be expected to
have a pure TiO2 stoichiometry. The O1s spectrum of group
4 was deconvoluted to O2, (PO4, CO3 and OH), and H2O,9
and showed no TiO2 sub-peak. This unique deconvolution

present only in this group is probably due to HA-blasting


method performed by the manufacturer; the surface was
completely covered by HA.
Some authors have used the difference between
the binding energy (BE) of an O1s sub-peak and the
corresponding Ti2p 3/2 peak to identify chemical states.
TiO2 has been referenced with a value of 71.5eV29. The

380

Chrcanovic et al.

present results from the samples of all groups corresponded


to this value.
Deconvolution of high resolution spectra of the carbon
1s peaks revealed three components for each surface: C=O,
C-O or C-OH, and C-C (graphite) or CH (hydrocarbon)6,29.
The binding energies for these three components in the
present study were ~288.53, ~286.80, and ~285.00eV,
respectively. The concentration of the hydrocarbon layer
was abruptly decreased after Ar sputtering, which indicates
that its source probably was due to adsorption of carbon
compounds from the ambient air, and was mainly located on
the oxide surface. This degree of contamination is conformal
to the percentages found after a common process of surface
cleaning32.

4.5. XRD analysis


The absence of peaks of Ca and P phases in the X-ray
diffractograms of the implants in groups 2 to 5, was
observed, while the XPS analysis showed the presence
of these elements. This happened because the layer of
Ca and P implants formed in groups 2, 3, and 5 is very
thin as to not be detected by XRD technique. In the case
of a very thin layer on the surface, the XRD would show
only the signal of the substrate while the XPS would be
very sensitive to the surface chemical composition (few
nanometers deep). In the case of group 4, Ca and P were
present in the hydroxyapatite. The results of XRD and XPS
suggest that the coating layer in implants from group 4 is
thicker than the other groups.
A coating process was developed in which HA was
blasted at a surface using standard sand blasting equipment,
and this was found to produce a HA coating which showed
promising results invivo33. This was the method used
(announced) by the implant manufacturer of the group 4
to incorporate HA on titanium surface. The fact that the
bioactive agent can be deposited unaltered suggests that the
technique could be used to deposit active therapeutic and
biologic agents33. The method used by the manufacturer
of group 4 implants (HA-blasting) was efficient for this
purpose.
It has been reported that the etching process modifies
the Ti surface composition of SLA-treated implants, and
X-ray diffraction (XRD) and metallographic microscopy
analysis indicated the presence of 20 to 40% of titanium
hydride (TiHx, x2) in addition to Ti.34 As the surface of
the implants in group 1 were also modified by double acid
etching as the SLA-treated implants, it is not surprising to
find titanium hydride in the implants from this group.
In one study35, TiO2 layers were treated by annealing
for 2hours in air at 600 and 1000C, which produce the
anatase and rutile phases, respectively. The quasi-amorphous
oxides can be crystallized by heating, which explains why
anatase was found in the group 2 of implants, since it was
subjected to an anodizing process. The manufacturer of
this group has not disclosed, but an annealing is done after
anodizing in order to obtain the anatase phase, because
it has some biological advantages: the anatase phase of
titania film enhances osteoblast adhesion, proliferation and
differentiation by affecting surface contact angles and/or
wettability36.

Materials Research

According to recent review made by Junkeretal.37,


there is sufficient proof that surface roughening induces a
safe and predictable implant-to-bone response, but it is not
clear whether this effect is due to the surface roughness
or to the related change in the surface composition. The
review of the experimental surface alterations revealed
that thin CaP coating technology can solve the problems
associated with thick CaP coatings, while they still improve
implant bone integration compared with non-coated
titanium implants. Moreover, it has been documented
that HA-coated implants have higher osseoconductive
properties when compared with uncoated implants38,39,
and even when compared with CaP-coated implants39.
In one study 38, primary human osteoblast-like cells
incubated on the HA-coated titanium appeared to exhibit
the highest number of cell process attaching to the surface
in comparison with three other different titanium surfaces
(as-machined, Al2O3-blasted, and plasma-sprayed with
titanium particles), which was indicative of optimal cell
attachment on HA. Considering this set of information,
the results presented here suggest the group 4 of implants
(HA-coated dental implants) may have the more suitable
characteristics for faster osseointegration.

5. Conclusions
It was confirmed considering the experimental results
present here, that the titanium with different coatings show
a wide range of chemical, physical properties, and surface
topographies or morphologies, depending on how they
are prepared and handled. Furthermore, by using different
surface preparation methods it is possible to control and vary
specific surface properties of titanium over a relatively wide
range. Different chemical properties of protective layer of
titanium-based implants analyzed in this study show different
behavior of deposited layer. The XPS analysis showed that
the surface of the implants studied were mainly composed
of a layer of TiO2, but there are other oxides associated with
a lesser amount as TiO and Ti2O3. Organic contamination
was observed on all surfaces which is not surprising as
every titanium (oxide) surface exposed to air will adsorb
hydrocarbons or carbon-oxygen containing species from
air. Different concentrations of calcium and phosphor were
measured on the surface of some samples, because this kind
of characterization is used by each manufacturer in order to
produce a bone conductor substrate. From XRD results in the
present study, the HA-blasting method used (announced) by
the implant manufacturer of the group 4 to incorporate HA
on titanium surface seems to be an efficient method for this
purpose.
According to the manufacturers, the addition of Ca
and P at the implant surface is a main feature of these
implants (groups 2 to 5). However, the results showed a
great discrepancy on the final amount of these elements
on the implant surface, which suggests a different
effectiveness of the employed surface finishing methods to
fix these elements on the implant surface tested here. All
manufacturers of this study that tried to incorporate Ca and
P on the titanium implant surface (groups 2, 3, 4 and 5)
succeeded, according to the results presented by the EDS.

2012; 15(3)

Chemical and Topographic Analysis of Treated Surfaces of Five Different Commercial Dental Titanium Implants

But only the manufacturer of the group 4 implants could


effectively incorporate HA on titanium. The other groups
(2, anodization and incorporation of Ca/P; 3, acid etching
and deposition of Ca/P; and 5, Ca/P-blasting and acid
etching) yielded only the formation of superficial amorphous
Ca/P complex. The implants from group 4 also showed the
highest roughness parameters.

381

Acknowledgements
This work was supported by the Brazilian agencies
CNPq and FAPEMIG. The authors would like to thank
Sai Sunil Kumar, Renato de Mendona, den Cristiano
Costa, Mrio da Silva Arajo Filho, and Antnio Luis Neto
Custdio.

References
1. Orsini G, Piattelli M, Scarano A, Petrone G, Kenealy
J, Piattelli Aetal. Randomized, controlled histologic
and histomorphometric evaluation of implants with
nanometer-scale calcium phosphate added to the dual
acid-etched surface in the human posterior maxilla. Journal
of Periodontology.2007;78(2):209-218. PMid:17274708.
http://dx.doi.org/10.1902/jop.2007.060297
2. Albrektsson T and Johansson C. Osteoinduction,
osteoconduction and osseointegration. European Spine
Journal.2001;10(Suppl2):S96-S101. PMid:11716023. http://
dx.doi.org/10.1007/s005860100282
3. Stanford CM. Surface modification of biomedical and
dental implants and the processes of inflammation, wound
healing and bone formation. International Journal of
Molecular Sciences.2010;11(1):354-369. PMid:20162020.
PMCid:2821008. http://dx.doi.org/10.3390/ijms11010354
4. Wennerberg A and Albrektsson T. Effects of titanium
surface topography on bone integration: a systematic review.
Clinical Oral Implants Research.2009;20(Suppl4):172-184.
PMid:19663964. http://dx.doi.org/10.1111/j.16000501.2009.01775.x
5. Albrektsson T, Brnemark PI, Hansson H and Lindstm J.
Osseointegrated titanium implants. Requirements for ensuring
a long-lasting, direct bone-to-implant anchorage in man. Acta
Orthopaedica Scandinavica.1981;52(2):155-170. http://
dx.doi.org/10.3109/17453678108991776
6. Lausmaa J. Surface spectroscopic characterization of titanium
implant materials. Journal of Electron Spectroscopy and
Related Phenomena.1996;81(3):343-361. http://dx.doi.
org/10.1016/0368-2048(95)02530-8
7. Moulder JF, Stickle WF, Sobol PE and Bomben KD.
H a n d b o o k o f X - ra y P h o t o e l e c t ro n S p e c t ro s c o p y.
Minnesota: Perkin-Elmer Corporation, Physical Electronics
Division;1992.
8. Cullity BD and Stock SR. Chapter10: The Determination
of Crystal Structure. In: Cullity BD, Stock SR. Elements
of X-Ray Diffraction.3thed. Reading: Addison-Wesley
Publishing;1956. p.297-323.
9. Kaiulis S, Mattogno G, Pandolfi L, Cavalli, M, Gnappi
G and Montenero A. XPS study of apatite-based
coatings prepared by sol-gel technique. Applied Surface
Science.1999;151(5):1-5.
10. Castilho GAA, Martins MD and Macedo WAA. Surface
characterization of titanium based dental implants. Brazilian
Journal of Physics.2006;36(3b):1004-1008. http://dx.doi.
org/10.1590/S0103-97332006000600055
11. Massaro C, Rotolo P, De Riccardis F, Milell E, Napoli A,
Wieland Metal. Comparative investigation of the surface
properties of commercial titanium dental implants. Part I:
chemical composition. Journal of Materials Science: Materials
in Medicine.2002;13(6):535-548. PMid:15348583. http://
dx.doi.org/10.1023/A:1015170625506

12. Peters M, Hemptenmacher J, Kumpfert J and Leyens C.


Chapter1. Structure and Properties of Titanium and Titanium
Alloys. In: Leyens C, Peters M, editors. Titanium and
Titanium Alloys: Fundamentals and Applications. Weinheim:
Wiley-VCH;2003. p.1-36.
13. Li DH, Lin J, Lin DY and Wang XX. Synthesized
silicon-substituted hydroxyapatite coating on titanium
substrate by electrochemical deposition. Journal of Materials
Science: Materials in Medicine.2011;22(5):1205-1211.
PMid:21465241. http://dx.doi.org/10.1007/s10856-0114310-y
14. Kim KJ and Park YR. Structural and optical properties of
rutile and anatase TiO2 thin films: effects of Co doping. Thin
Solid Films.2005;484(1-2):34-38. http://dx.doi.org/10.1016/j.
tsf.2005.01.039
15. Kovalev DY, Prokudina VK, Ratnikov VI and Ponomarev
VI. Thermal Decomposition of TiH 2: A TRXRD Study.
International Journal of Self-Propagating High-Temperature
Synthesis.2010;19(4):253-257. http://dx.doi.org/10.3103/
S1061386210040047
16. Aparicio C, Padrs A and Gil FJ. In vivo evaluation of
micro-rough and bioactive titanium dental implants using
histometry and pull-out tests. Journal of Mechanical Behaviour of
Biomedical Materials.2011;4(8):1672-1682. PMid:22098868.
http://dx.doi.org/10.1016/j.jmbbm.2011.05.005
17. Wennerberg A. On roughness and implant incorporation.
[Tese]. Gteborg: Gteborg University;1996.
18. Wennerberg A, Albrektsson T and Lausmaa J. Torque and
histomorphometric evaluation of c.p. titanium screws blasted with
25- and 75-m-sized particles of Al2O3. Journal of Biomedical
Materials Research.1996;30(2):251260. http://dx.doi.
org/10.1002/(SICI)1097-4636(199602)30:2%3C251::AIDJBM16%3E3.0.CO;2-P
19. Wennerberg A, Hallgren C, Johansson C and Danelli S. A
histomorphometric evaluation of screw shaped implants each
prepared with two surface roughness. Clinical Oral Implants
Research.1998;9(1):11-19. PMid:9590940. http://dx.doi.
org/10.1034/j.1600-0501.1998.090102.x
20. Gotfredsen K, Berglundh T and Lindhe J. Anchorage of
titanium implants with different surface characteristics: an
experimental study in rabbits. Clinical Implant Dentistry and
Related Research.2000;2(3):120-128. PMid:11359256. http://
dx.doi.org/10.1111/j.1708-8208.2000.tb00002.x
21. Wrbel E, Witkowska-Zimny M and Przybylski J. Biological
mechanisms of implant osseointegration. Ortopedia Traumatologia
Rehabilitacja.2010;12(5):401-409. PMid:21057147.
22. Olefjord I and Hansson S. Surface analysis of four dental
implant systems. The International Journal of Oral and
Maxillofacial Implants.1993;8(1):32-40. PMid:8468084.
23. Sawase T, Hai K, Yoshida K, Baba K, Hatada R and Atsuta
M. Spectroscopic studies of three osseointegrated implants.

382

Chrcanovic et al.

Journal of Dentistry.1998;26(2):119-124. http://dx.doi.


org/10.1016/S0300-5712(96)00080-2
24. Sul YT, Johansson C, Chang BS, Byon ES and Jeong Y.
Bone tissue responses to Mg-incorporated oxidized implants
and machine-turned implants in the rabbit femur. Journal of
Applied Biomaterials and Biomechanics.2005;3(1):18-28.
PMid:20799236.
25. Keller JC, Draughn RA, Wightman JP, Dougherty WJ and
Meletiou SD. Characterization of sterilized CP titanium implant
surfaces. The International Journal of Oral Maxillofacial
Implants.1990;5(4):360-367. PMid:2094654.
26. Zitter H and Plenk HJ. The electrochemical behaviour of metallic
implant materials as indicator of their biocompatibility. Journal
of Biomedical Materials Research.1987;21(7):881-896.
PMid:3611146. http://dx.doi.org/10.1002/jbm.820210705
27. Solar RJ, Pollack SR and Korostoff E. In vitro corrosion
testing of titanium surgical implant alloys: an approach
to understanding titanium release from implants. Journal
of Biomedical Materials Research.1979;13(2):217-250.
PMid:429392. http://dx.doi.org/10.1002/jbm.820130206
28. Kasemo B and Lausmaa J. Biomaterial and implant
surfaces: on the role of cleanliness, contamination, and
preparation procedures. Journal of Biomedical Materials
Research.1988;22(A2 Suppl):145-158.
29. Ong JL, Lucas LC, Raikar GN and Gregory JC. Electrochemical
corrosion analyses and characterization of surface modified
titanium. Applied Surface Science.1993;72(1):7-13. http://
dx.doi.org/10.1016/0169-4332(93)90036-B
30. Carley AF, Chalker PR, Riviere JC and Roberts MW. The
identification and characterization of mixed oxidation states
at oxidised titanium surfaces. Journal of Chemical Society,
Faraday Transactions1: Physical Chemistry in Condensed
Phases.1987;83(2):351-370.
31. Sham TK and Lazarus MS. X-ray photoelectron spectroscopy
(XPS) studies of clean and hydrated TiO2 (rutile) surfaces.
Chemical Physics Letters.1979;68(2-3):426-432. http://dx.doi.
org/10.1016/0009-2614(79)87231-0
32. Cacciafesta P, Hallam KR, Oyedepo CA, Humphris
ADL, Mervyn JM and Jandt KD. Characterization
of ultraflat titanium oxide surfaces. Chemistry of
Materials.2002;14(2):777-789.

Materials Research

33. OHare P, Meenan BJ, Burke GA, Byrne G, Dowling D and


Hunt JA. Biological responses to hydroxyapatite surfaces
deposited via a co-incident microblasting technique.
Biomaterials.2010;31(3):515-522. PMid:19864018. http://
dx.doi.org/10.1016/j.biomaterials.2009.09.067
34. Perrin D, Szmukler-Moncler S, Echikou C, Pointaire P
and Bernard JP. Bone response to alteration of surface
topography and surface composition of sandblasted and
acid etched (SLA) implants. Clinical Oral Implants
Research.2002;13(5):465-469. PMid:12453122. http://dx.doi.
org/10.1034/j.1600-0501.2002.130504.x
35. Nam SH, Choi JW, Cho SJ, Kimt KS and Boo JH. Growth of
TiO2 anti-reflection layer on textured Si (100) wafer substrate
by metal-organic chemical vapor deposition method. Journal
of Nanoscience and Nanotechnology.2011;11(8):73157318. PMid:22103185. http://dx.doi.org/10.1166/
jnn.2011.4813
36. He J, Zhou W, Zhou X, Zhong X, Zhang X, Wan Petal.
The anatase phase of nanotopography titania plays
an important role on osteoblast cell morphology and
proliferation. Journal of Materials Science: Materials in
Medicine.2008;19(11):3465-3472. PMid:18592349. http://
dx.doi.org/10.1007/s10856-008-3505-3
37. Junker R, Dimakis A, Thoneick M and Jansen JA. Effects
of implant surface coatings and composition on bone
integration: a systematic review. Clinical Oral Implants
Research.2009;20(Suppl4):185-206. PMid:19663965. http://
dx.doi.org/10.1111/j.1600-0501.2009.01777.x
38. Prado Da Silva MH, Soares GD, Elias CN, Best SM,
Gibson IR, DiSilvio Letal. In vitro cellular response to
titanium electrochemically coated with hydroxyapatite
compared to titanium with three different levels of
surface roughness. Journal of Materials Science:
Materials in Medicine.2003;14(6):511-519. http://dx.doi.
org/10.1023/A:1023455913567
39. Yang GL, He FM, Hu JA, Wang XX and Zhao SF.
Biomechanical comparison of biomimetically and
electrochemically deposited hydroxyapatite-coated porous
titanium implants. Journal of Oral and Maxillofacial
Surgery.2010;68(2):420-427. PMid:20116717. http://dx.doi.
org/10.1016/j.joms.2009.09.014

También podría gustarte